Introns and Their Therapeutic Applicationsin Biomedical Researches

Document Type : Review Paper

Authors

1 Industrial Biotechnology Research Group, Institute of Biotechnology, Ferdowsi University of Mashhad, Mashhad, Iran

2 Stem Cell and Regenerative Medicine Research Group, Iranian Academic Center for Education, Culture and Research (ACECR), Khorasan Razavi Branch, Mashhad, Iran

3 Biotechnology Research Center, Pharmaceutical Technology Institute, Mashhad University of Medical Sciences, Mashhad, Iran

4 Applied Biomedical Research Center, Mashhad University of Medical Sciences, Mashhad, Iran

5 Department of Biology, Faculty of Science, Ferdowsi University of Mashhad, Mashhad, Iran

6 National Institute for Genetic Engineering and Biotechnology, Tehran, Iran

Abstract

ntext: Although for a long time, it was thought that intervening sequences (introns) were junk DNA without any function, their critical roles and the underlying molecular mechanisms in genome regulation have only recently come to light. 
Introns not only carry information for splicing, but they also play many supportive roles in gene regulation at different 
levels. They are supposed to function as useful tools in various biological processes, particularly in the diagnosis and 
treatment of diseases. Introns can contribute to numerous biological processes, including gene silencing, gene imprinting, 
transcription, mRNA metabolism, mRNA nuclear export, mRNA localization, mRNA surveillance, RNA editing, NMD, 
translation, protein stability, ribosome biogenesis, cell growth, embryonic development, apoptosis, molecular evolution, 
genome expansion, and proteome diversity through various mechanisms.
Evidence Acquisition: In order to fulfill the objectives of this study, the following databases were searched: Medline, Scopus, Web of Science, EBSCO, Open Access Journals, and Google Scholar. Only articles published in English were included.
Results & Conclusions: The intervening sequences of eukaryotic genes have critical functions in genome regulation, 
as well as in molecular evolution. Here, we summarize recent advances in our understanding of how introns influence 
genome regulation, as well as their effects on molecular evolution. Moreover, therapeutic strategies based on intron sequences are discussed. According to the obtained results, a thorough understanding of intron functional mechanisms could 
lead to new opportunities in disease diagnosis and therapies, as well as in biotechnology applications.

Keywords

Main Subjects


1. Background

The first use of DNA sequencing in the late 1970s revealed that genes in eukaryotic cells are discontinues, i.e. they have an intron /exon structure ( 1 ). More DNA sequencing has revealed that introns make up the majority of sequences in most eukaryotic genes. However, it has also been demonstrated that these regions are removed after transcription, leading to the conclusion that they are unnecessary parts of DNA ( 2 , 3 ). These features raise some questions: if introns are considered junk, what is the reason for their existence and why haven’t they been eliminated through evolution? Moreover, it is known that eukaryotic cells expend significant amounts of energy to enhance and maintain intron sequences. As a result, a debate about the function of introns emerged shortly after their discovery ( 1 , 4 ) . So, the early functions related to introns include the simplified development of unique genes and the possibility for the expression of various proteins from a single gene ( 5 ). Additionally, introns play a critical role in protecting coding regions in eukaryotic genomes from frequent errors ( 6 ). Gradually, more experimental analyses revealed that introns or their derived elements have significant effects on gene expression in a wide range of organisms, including insects, nematodes, and mammals ( 7 ). However, their exact mechanisms are still not fully understood. For example, several reports, have demonstrated that certain introns are crucial for the accumulation of specific mRNAs such as rat growth hormone and growth hormone receptor ( 8 , 9 ), triosephosphate isomerase ( 10 ), and human β-globin ( 11 ). Moreover, evidence in plants shows that introns act post-transcriptionally to enhance mRNA processing, possibly, by facilitating mRNA maturation or increasing the stability of nascent transcripts ( 12 , 13 ). These reports indicate that introns and splicing are necessary for mRNA to be transported from the nucleus. However, analyzing the structure of certain genes, such as histones and olfactory receptors, in higher vertebrates has shown that the presence of introns and splicing is not necessary for the functioning of all genes ( 1 ). On the other hand, the absence of introns in some genes can result in both reduction and repression of gene expression ( 14 , 15 ), but it does not have any effect on other genes ( 1 ).

Nevertheless, while accepting that gene expression is affected by certain intron sequences, there is still a debate about the mechanisms behind this that needs to be elucidated In addition, analyzing the structure of genes has revealed that many of them contain 10-50 introns ( 1 , 16 ). Consequently, this raises further questions, such as whether all introns of a gene have an impact on gene regulation. Do they act with the same mechanism? Meanwhile, which intron has a greater impact on the regulation of the corresponding gene? To answer these questions, many investigations have been conducted on factors affecting intron activities. These investigations have found that intron sequence context (ISC) and intron position ( 17 - 22 ) play a significant role. In addition, it has become increasingly clear that intron excision by the spliceosome can affect multiple steps of mRNA metabolism. These steps include primary transcription of the gene, correction and polyadenylation of the pre-mRNA, nuclear export, translation, and degradation of the mRNA product ( 5 , 23 , 24 ). Exon and intron sequence context not only play critical roles in gene regulation but also have essential functions in molecular evolution ( 1 , 23 , 25 ). In this review, we present a brief and up-to-date overview of how introns exert their effects on eukaryotic gene regulation and molecular evolution. Additionally, their application in diagnosis and treatment will be discussed.

2. Splicing and Splicing Control Elements

Precise removal of introns by splicing (Fig. 1) before mRNA maturation is a crucial and universal step in higher eukaryotic genes. This process yields suitable mRNAs for gene expression ( 26 , 27 ). Although the accuracy and complexity of intron removal still amazes even 30 years after the discovery of introns ( 28 ), it has been determined that intron excision and splicing includes several step by-step assembly and catalytic processes. These processes comprise exon and intron diagnosis, intron cutting, and exon joining ( 29 , 30 ), which are controlled by splicing control elements: 1. Classical cis-controlling elements, are weakly conserved intronic cis-elements, which are essential for defining the exon boundaries, including GU and AG dinucleotides at the exon–intron and intron–exon junctions, respectively (5’- and 3’-splice locations), an A nucleotide at the branch point and a polypyrimidine tract (Py)n with variable length upstream of the 3’-splice site. The branch point is usually placed 18–40 nucleotides upstream from the polypyrimidine tract. 2. Basal splicing machinery is an enormous complex macromolecule that is composed of as many as 300 distinct proteins acting as RNA recognition motifs (RRMs) or ATPases and five RNAs (snRNAs) (U1, U2, U4, U5 and U6) which its components bind to classical cis-controlling elements and stimulate gathering of the splicing complex. 3. Cis auxiliary elements are exonic and intronic cis-elements, which are generally necessary for effectual splicing of constitutive and alternative splicing. 4. Trans auxiliary elements can interact with enhancers and silencers ( 31 - 34 ). These elements interact with the basal splicing machinery, specifically the spliceosome complex (Fig. 1), which acts as a trans-acting element ( 32 , 35 ). However, auxiliary elements are additional information that act as auxiliary cis-acting elements, such as exonic/intronic splicing enhancers/silencers (ESE/S, ISE/S), and auxiliary trans-acting factors, such as serine arginine-rich (SR) proteins and heterogeneous nuclear ribonucleoproteins (hnRNPs).

Figure 1.Splicing and controlling splicing elements.A) Schematic representation of splicing which is performed as a co-transcriptional process. B) Splicing control elements. Classical cis-regulatory elements are indicated by blue letters and boxes. ESE: exon splicing enhancer; ESS: exon splicing silencer; ISS: intron splicing enhancer; ISS: intron splicing silencer; SR: serine arginine; hnRNPs: heterogeneous nuclear ribonucleoprotein particles.

Overall, SR family binding at ESE and ISE simplifies exon recognition although hnRNPs are inhibitory. These elements contribute to accurate splicing by aiding in the identification of suitable splice sites or by suppressing or increasing the usage of certain splice sites, particularly in alternative splicing ( 31 , 32 ) ( 36 , 37 ).

2.1. Effects of Splicing Control Elements (SCEs) On Gene Regulation

Regarding, the contribution of SCEs in the regulation of gene expression, it has been determined that various genetic diseases such as hereditary breast and ovarian cancer (caused by the c.4185 + 4105C > T variant in BRCA1, the first reported deep intronic variant in this gene that activates a pre-existing cryptic donor site) and Ataxia-telangiectasia disease (due to the c.2839-581_2839-578del variant in the ATM gene, which creates an ISE), can be resulted from SCEs malfunctioning ( 23 ) ( 38 - 40 ). Among spliceosome-associated catalytic entities (SCEs), the spliceosome plays a crucial role in splicing and gene expression regulation, particularly at the transcriptional level ( 14 , 41 ). Additionally, in addition to its role in RNA processing, the components of the spliceosome complex interact with the TFIIH subunit of RNA polymerase II. This subunit regulates various stages of transcription, both in vivo and in vitro, including abortive initiation, re-initiation, promoter escape, and early elongation (Table 1) ( 42 - 47 ). More-over, it has been demonstrated that some SCEs can also affect the mRNA nuclear export (Table 1) ( 48 - 50 ).

Pathways Regulatory elements Activities Mechanisms Regulation levels Sample of biological processes Reference
SCEs U1 Transcription Interaction with TFIIH Transcription and Post transcription Embryonic development, Cell growth, Apoptosis ( 31 , 32 , 41 , 45 , 179 )
Splicing
Other elements Splicing Interaction with SCEs
alternative splicing
mRNA nuclear transport Interaction with nuclear transport
EJCs Upf2, Upf3 Y14, RNPS1, Magoh, eIF4AIII, MLN51 RNPS1, ACINUS NMD Distinction between PTC and NTC Translation mRNA quality control ( 49 , 53 , 60 , 67 , 138 , 180 , 181 )
Y14; Magoh mRNA localization Imprints some of the information Post transcription Regulation of sub-cytoplasmic regions
Yap56; Aly; REF SRm160 Tap-p15; PININ mRNA nuclear transport Splicing-coupled mechanisms Post transcription Overcome the nuclear retention
IDREs SnoRNAs and scaRNAs rRNA modifications Guide for nucleoside modifications Translation Ribosome biogenesis ( 66 , 81 , 85 , 182 )
snRNA modifications Post transcription Splicing
tRNA modifications Transcription Translation
mRNA modifications Post transcription mRNA transport
Orphan ? ? ?
MiRNAs Cleared undesired mRNA Inhibition of transcription Transcription Cell transitions quickly, Developmental timing, Apoptosis and tissue growth ( 68 , 69 , 76 )
Suppression of translation Translation
? ?
nmRNAs Regulate HLTFG genes As activator Transcription Developmental stage ( 88 , 90 , 91 , 111 )
TFs Transcription Factor Contribute in transcription Transcription Numerous ( 41 , 183 )
IREs Providing regulatory motifs Transposition Transcription Molecular evolution, RNA editing, Splicing and apoptosis, Cell growth and differentiation ( 41 , 98 , 126 , 171 )
Post transcription
Translation
?
ISCS PAs Contribution to polyadenylation Polyadenylation coupled mechanisms mRNA Transport, mRNA Survival, mRNA stability mRNA Transport ( 18 , 41 , 184 )
Translation
Post translation
TFBs Providing sites for TFs Contribute in transcription Transcription Numerous ( 23 , 41 , 183 )
CpG island Contribution to transcription Transcription Transcription Numerous ( 23 , 41 , 183 )
Enhancer RNA polymerase assembly Increase transcription Transcription Numerous ( 23 , 41 , 183 )
Promoter RNA polymerase assembly Transcription Transcription Numerous ( 23 , 41 , 183 )
GR Targets for gene regulation Contribute in transcription Contribute in polyadenylation Transcription Gene expression, Genome stability, Evolution ( 20 , 95 , 108 , 109 )
Post transcription
Translation
Table 1.Possible mechanisms by which intron can regulate expression of genes

2.2. Splicing Effects on the Gene Expression Regulation

In addition to direct effects of SCEs on gene regulation, it has been demonstrated that SCEs also play a role during the splicing process. During splicing, a protein compound called the Exon Junction Complex (EJC), which consists of a stable heterotetramer core and other factors, is placed 20 to 24 nucleotides upstream of each exon-exon junction by the splicing machinery (Fig. 2) ( 51 - 54 ). This complex plays a crucial role in various processes related to mRNA, including its stability, survival, transport, and translation ( 23 , 55 - 57 ).

Figure 2.Formation, components and effects of exon junction complexes.A) EJC formation. B) EJC composition. C) A model of mRNA nuclear transport. D) Motor-based active mRNA localization. E) Translation or NMD. PTCs are typically followed by minimum one EJC (curved arrows in E). Motor protein needed in mRNA nuclear transport is indicated by green triangle. MNT: mRNA nucleolar transport; ML: mRNA localization; NMD: nonsense mediated mRNA decay.

It is involved in post-transcriptional processes such as mRNA nuclear transport and translational regulation. The complex is formed through a splicing-coupled model, where components of EJC bind to mRNA and nuclear pore elements. EJCs are made of three types of proteins based on their functions, including MNT, ML and NMD proteins, which are involved in mRNA nucleolar transport, mRNA localization and nonsense mediated mRNA decay, respectively. These proteins based on their function have different domains such as RNA-binding domain (RBD), nuclear transport domain (NTD) and mRNA localization domain (MLD) and after doing their function, they dissociate from EJC (Fig. 2) ( 49 , 52 , 58 - 60 ). Additionally, it is responsible for mRNA localization using imprinting information, which is necessary for proper cytoplasmic localization. Splicing-coupled model, mRNA transport using the channel needs binding of the mRNA to solvable transport receptors, which include MNT proteins of the EJC in this model. These receptors mediate transport in an energy independent manner by interacting with the phenylalanine-glycine repeats, a group of nucleoporins that line the channel of the nuclear pore complex, and detach as of mRNP after transfer to the cytoplasm and other components of EJCs, which are involved in ML and NMD remain attached to the mRNP. For additional details, refer to Table 1. Active transport of mRNAs from the nucleus to peripheral cytoplasmic target positions, needs ML factors, which act as a memory and imprint through splicing. In this model, mRNP particles movement along cytoskeletal filaments is associated with a motor protein, most probable with adaptor proteins, and is transported to the aim locate as dormant. Finally, at destination sites the ML factors and other RNPs such as those that repress translation during transport are released from the mRNA and are fastened to the select locate using particular proteins and recruit ribosomes and other translational components (Fig. 2) ( 23 , 61 - 63 ). The complex also participates in nonsense-mediated mRNA decay (NMD), mRNA quality control, and translation by distinguishing between premature and normal termination stop codons (PTC and NTC) (Fig. 2D) ( 23 , 53 , 64 - 67 ).

2.3. Intronic-Derived Regulatory Elements

Major effects of the introns on gene regulation and molecular evolution, whether direct or indirect, are associated with regulatory elements that arise from intronic sequences. These elements are involved in the regulation of transcription or post-transcriptional processes. MicroRNAs (miRNAs) act as trans-vehicles in gene regulation, preventing transcription and/or repressing translation ( 68 - 72 ). They can be derived from intronic regions. These elements, which are diverse in structure and function, are widely distributed and abundant in eukaryotic cells. They are divided into four classes based on their origins, including intergenic (In), intronic (Id), palindromic (P), and exonic (E). Id-miRNAs are the only class of miRNAs that are derived from introns ( 68 , 69 , 73 - 75 ). Id-miRNAs are transcribed by the Pol-II promoters of the encoding genes and are co-expressed in the intron areas of the gene transcriptions (pre-miRNA). After RNA splicing and additional treating, the spliced intron might function as a pri-miRNA for Id-miRNA generation. In the nucleus, the pri-miRNA is excised by Drosha RNase to create a hairpin-like pre-miRNA pattern, which is later transferred to the cytoplasm for more treating by Dicer to make mature and active miRNA (Fig. 3B) ( 76 ). The Id-miRNAs which can be considered gene expression products, have the ability to interfere with the expression of other genes ( 76 , 77 ). Hence, it has been suggested that these elements are capable of rapidly triggering cell transitions in response to external stimuli without the need for time-consuming protein synthesis ( 69 ). They also play critical roles in various biological functions, including developmental timing, apoptosis, and tissue development ( 76 , 78 , 79 ). Moreover, small nucleolar RNAs (snoRNAs) are another type of intronic-derived regulatory elements (IDREs) that originate only from intronic sequences. They are located in the nucleoli and act as cofactors in ribosome biogenesis (Fig. 3C) ( 66 , 80 - 83 ). Meanwhile, the roles of snoRNAs are not restricted to ribosome synthesis. They are also used as guides for modifying other cellular RNAs, including snRNAs, tRNAs, and mRNAs ( 66 , 81 , 84 ). Nonetheless, the functions of a large number of these “orphan” snoRNAs remain unknown, and there are likely even more snoRNAs waiting to be discovered. In wholly stated cases, vertebrate snoRNAs initiate from introns of either protein coding/noncoding RNA polymerase II-transcribed genes. Here the pathways of snoRNA derivation from protein-coding gene are demonstrated. The main pathway is similar to id-miRNA biogenesis, whereby after splicing lariat undergoes several processing including lariat debranching, 5΄ and 3΄ exonucleation olising; at continue mature snoRNA assembly by accessory proteins and then transported from the nucleoplasm into the nucleolus as snoRNP complex. In the other pathway snoRNA is produced by endonucleolytic processing as directly from pre-mRNA ( 66 , 85 ). Nuclear messenger RNAs (nmRNAs), a group of small non-coding RNAs (sncRNAs), are involved in the developmental program by regulating the expression of high-level transcription factor genes (HLTFGs) such as Hox genes, as well as some other protein-coding genes. These nmRNAs, along with certain types of IDREs, function similarly to miRNAs and snoRNAs ( 86 - 92 ). Interestingly, the introns of these genes may represent elements of the generation-specific control keys (GSCK) that act as non-coding mRNA (Fig. 3D) ( 91 ). Other types of sncRNAs, such as small cajal body-specific RNAs (scaRNAs), also originated from introns ( 93 ). Perhaps many other types of these IDREs remain to be discovered. For additional details, refer to Table 1.

Figure 3.Schematic diagram of biogenesis of the intronic derived regulatory elements.A) Transcription and splicing processes. B) Biogenesis of Id-miRNA. C) snoRNA synthesis. D) nmRNA biogenesis. E) Intronic repetitive elements. The effects of introns on gene regulation and molecular evolution might be the result of intronic repetitive elements (IREs) by providing regulatory motifs throughout mobilization as replicative and conservative forms.

2.4. Sequence Context and Structure of the Introns asRegulatory Elements

Major sections of the intronic regions of the eukaryotic genome are tightly regulated, which can be attributed to the presence of regulatory elements in these regions ( 94 ). In this context, changes in certain sequences, including the intron regions, may result in biological malfunction ( 68 , 69 ). Apart from trans-acting regulatory elements derived from introns, many regulatory functions are attributed to the intron sequence context and structures (ISCSs), which are thought to function as cis-acting regulatory elements ( 20 , 95 ). The intron sequence context can contribute to gene regulation through various stages of mRNA-related processes. It provides regulatory elements such as promoter- and enhancer-like structures (Fig. 4 D), transcription factor binding sites (Fig. 4E), CpG islands (Fig. 4F) as well as splicing and polyadenylation sites ( 14 , 41 , 44 , 96 - 99 ).

While intron remains mainly in the double-stranded (Watson-Crick base pair) form, it can also form other complex structures, such as wobble base pairs, Hoogsteen triplexes, and G-quadruplexes. These structures are distributed non-randomly and may contribute to diverse biological activities during gene expression, as well as in genome stability and evolution ( 20 , 95 , 100 - 102 ). Nonetheless, very little is known about how their non-randomness contributes to biological functions. In the middle of the most interesting non-random sequences of the genome are G-rich sections (Fig. 4B), which have the potential to form G-quadruplex DNA or G-tetraplex DNA (Fig. 4C) ( 20 , 103 - 107 ). G-quadruplex DNA provides targets for the regulation of gene expression by binding agents that are crucial for transcription or splicing ( 20 ). These structures are distributed throughout both intra- and inter-regions of genes, with a positional bias towards the 5’- and 3’-ends of the gene. This bias suggests a specific function, particularly in the initiation and termination of gene transcription, respectively ( 20 , 95 ). In humanoid genomics, the number of locations with potential for the formation of G-quadruplex structures is estimated to exceed 300,000 ( 108 , 109 ). Some of the potential G-quadruplex regions are located within introns and are correlated with the function of the corresponding protein. In particular, proto-oncogenes are rich in G nucleotides, while tumor suppressor genes have a lower frequency of G-runs compared to the genomic average ( 20 ). Promoter regions are also G-rich areas ( 110 - 113 ). Interestingly, positional biases of G-rich areas at the 5’-ends of the first introns of genes from frogs to humans have been established. These biases may provide structural targets for regulating gene expression at the transcription or RNA processing levels ( 20 , 100 - 102 ). In our previous research, we have provided evidence supporting the existence of promoter- and enhancer-like structures near CpG islands in the first intron of the human factor VIII gene ( 114 ).

Figure 4.Sequence context and structure of the introns.A) Schematic diagram of the complexity of the genome and a model of eukaryotic gene structure. B) G-rich regions (red box). C) Schematic diagram of a G-quadruplex with four guanines arranged around a central monovalent cation. D) Intronic and exonic promoter (blue line), these structures with variant LDF can be distributed as random throughout a gene including intron and exons, nonetheless some of them can have significant distribution in correlation with other regulatory motifs including CpG islands (green lines, E), TFBs (brown line, F), splicing and polyadenylation sites (violet lines, G) and REs (yellow area, H).

3. Introns and Gene Regulation Pathways

In general, introns and their associated processes can modulate any step of gene regulation. However, transcription, post-transcriptional modification, transla-tion, and post-translational modifications are the main steps in gene expression where gene regulation could occur. These steps can be directly and/or indirectly influenced by introns ( 41 , 45 , 115 , 116 ). Although many questions regarding the regulatory mechanisms and their related effector elements have been answered so far, a large number of them remain unknown and await discovery. In Table 1, a list of different gene regulation pathways is presented, in which either introns or intronic elements are involved. Some of the biological processes controlled by these pathways and elements are indicated.

3.1. Effects of Introns on Proteome Expansion

In higher eukaryotes, comparative analysis between genes and their corresponding expressed sequence forms indicates a highly intricate process of gene regulation. This process leads to the expression of different proteins with various and sometimes even antagonistic functions from a single gene ( 23 , 117 - 121 ). Alternative splicing, trans-splicing, and RNA editing are three major processes responsible for proteome complexity. These processes are modulated by cell type, developmental phase, gender, and/or in response to external stimuli ( 32 , 118 , 119 , 122 - 125 ). So, introns, which can be involved in alternative splicing, trans-splicing, and RNA editing, play significant roles in proteome complexity by affecting transcription, mRNA localization, mRNA stability, and translation ( 23 , 120 , 126 , 127 ). Moreover, RNA editing and splicing can also be influenced by intronic repetitive elements (IREs), which can result in transcriptome expansion ( 126 , 128 , 129 ). Altogether IREs influence transcriptional, post-transcriptional, and translational levels ( 98 , 126 , 130 ) by providing regulatory motifs throughout the process of mobilization ( 126 , 131 ).

3.2. The Effects of Introns on Molecular Evolution

Current advances in genomics, including whole-genome sequencing, high-throughput protein characterization, and bioinformatics, have led to a significant improvement in studies on molecular evolution. The evolution (birth, death, and maintenance) of introns in eukaryotic species is a highly debated topic with many unresolved questions. In an interesting study, Li et al. unravel some of the key aspects of intron evolution. They studied a species of microcrustaceans, Daphnia pulex, which has been shown to exhibit high intron dynamics. They sequenced and assembled nine complete individual genomes from distinct natural populations. Using a dedicated bioinformatics pipeline, they identified 90 recently gained introns. They reported contrasting conclusions compared to previous studies based on more ancient intron gains. In the species under study, they found that (i) intron gains are rather common and serve as an important source of genetic variation, (ii) intron gains are more frequent than intron losses, and (iii) parallel intron gains occur frequently. The authors reported that intron gains are, on average, slightly deleterious. All intron gains reported in this study result from the repair of double-strand breaks (DSBs), either a single blunt DSB or a single staggered DSB. The authors demonstrated that these mechanisms can lead to multiple rounds of double-strand break (DSB) repair, which can increase the length and complexity of introns. They also proposed the hypothesis that certain intron sequences may arise from the de novo synthesis of DNA caused by DNA polymerase errors ( 132 ).

Here, we emphasize the central role of introns in evolutionary processes. We argue that the effects of introns, specifically through IDREs (Intron-Derived Regulatory Elements) and ISCSs (Intron Splicing Control Sequences), could generate different sets of RNAs. These RNAs are then subject to natural selection based on the phenotypes they produce. IREs, as a type of IDREs and ISCSs ( 98 , 133 , 134 ), have critical roles in molecular evolution. They influence transcription, post-transcription, and translational levels of gene regulation by providing regulatory motifs throughout mobilization ( 98 , 126 ). Moreover, RNA editing and splicing, which lead to transcriptome expansion, can be influenced by IREs ( 126 , 128 , 135 ). The effects of introns on molecular evolution are not restricted to IREs; they are also influenced by RNA editing-based introns, which serve as a guide for modification ( 5 , 136 ). On the other hand, since intronic sequences are more variable than exons, these changes lead to the formation of new double-stranded RNA (dsRNA) editing substrates. Due to alternative splicing and RNA editing, the amino acid sequence of proteins can be altered, resulting in the creation of new phenotypes that may undergo natural selection. Thus, RNA editing based on introns plays an important role in molecular evolution ( 5 ). Besides ontogeny, which refers to the development of a zygote into a complex multi-cellular organism, the process can also be seen as a miniature of evolution and is influenced by introns ( 91 ). In this regard, although alternative splicing plays critical roles in developmental stages ( 120 , 137 ), recent reports have demonstrated that intron-derived regulatory elements, such as non-messenger RNAs (nmRNAs), also contribute to this process ( 91 ).

3.3. The Effects of Introns on Disease Development, Diagnosis, and Healing

Due to the assumed effects on gene regulation and its role in biological processes, any changes in introns and related elements may result in hereditary diseases. In this regard, several disorders, such as myotonic dystrophy, fragile X mental retardation, and dominant β-thalassemia, are caused by dysregulation of Id-miRNAs and NMD, respectively ( 51 , 69 , 138 ). So, this association can be used for both diagnostic and therapeutic purposes. In this regard, alternative splicing patterns of genes have been suggested as tools for diagnosis and treatment ( 139 , 140 ). The splicing design of a gene can be altered by both extracellular stimuli and mutations in splicing control elements ( 32 , 139 , 141 ). In contrast to extracellular stimuli such as hormones, immune response, neuronal depolarization, and cellular stress, which typically have temporary effects during normal development, splicing mutations often lead to hereditary diseases ( 119 , 120 , 142 ). Moreover, it has been shown that these mutations might be linked to cancer ( 119 , 143 - 145 ). In this regard, the alternative splicing of genes involved in apoptosis, angiogenesis, adhesion and metastasis, invasion, propagation, and hormone signaling is now well-recognized in cancer ( 119 , 144 , 146 , 147 ). For example, the CD44 gene comprises 10 variable exons (v1–v10) and expresses a family of hundreds of diverse CD44 isoforms. These isoforms are cell-surface glycoproteins that function in cell adhesion, migration, and matrix interactions through alternative splicing. This includes both standard and variable forms ( 139 , 148 ). Although the standard isoform of CD44 is predominant in healthy human adult tissues, variable isoforms are expressed in some tissues during development and T-cell activation ( 139 , 148 ). Meanwhile, a group of variable isoforms of CD44 with metastatic potential are expressed in numerous human malignancies ( 60 , 120 , 123 , 139 ). Moreover, several genes involved in apoptosis, such as Bcl-x and caspase 9, can express antagonistic pro- and anti-apoptotic isoforms, as identified by alternative splicing ( 32 , 149 - 151 ). Consequently, the patterns of alternative splicing of certain genes can provide a tool for determining cellular states, including developmental stages and disease ( 152 ). Based on this information, novel therapeutic strategies for correcting or circumventing splicing abnormalities are now emerging. These methodologies include over-expression of proteins that alter the splicing of the affected exons ( 153 - 155 ), the use of oligonucleotides to obstruct the usage of improper splice sites and force the use of functional splice sites ( 139 , 156 - 159 ), the use of compounds that influence the phosphorylation of splicing agents such as hnRNPs and SR proteins ( 139 , 160 , 161 ), or stabilize putative secondary structures ( 162 ), and a trans-splicing method to substitute mutated exons with wild-type exons ( 124 , 125 , 163 ).

3.4. Intron Applications in Gene and Protein Therapy

Two main approaches for managing genetic disorders are replacement therapy with recombinant proteins and gene therapy. To generate an effective protein, an efficient expression vector with appropriate regulatory factors is necessary for both systems ( 41 ). Based on the effects of introns and splicing processes on gene regulation, utilizing introns could be a valuable approach for modifying expression vectors with appropriate regulatory elements. In this regard, it has been stated that some mRNAs transcribed from cDNA failed to exit the nucleus and hence did not produce a protein. However, the same mRNAs expressed from intron-containing constructs were able to enter the cytoplasm and be effectively translated, even up to 500-fold ( 18 , 41 , 114 , 164 , 165 ). In our previous study, we introduced hBG intron-II in the second intronic position of the hFIXcDNA and observed an increase in the expression level compared to that of the intron-less hFIX-cDNA in vitro ( 41 ). Nonetheless, each regulator of spliceosome catalytic elements (SCEs), exon junction complexes (EJCs), intron-defined regulatory elements (IDREs), and intron splicing control sequences (ISCSs) has a specific function that modulates gene expression at different levels (Table 1). Therefore, harnessing this potential relies on having precise knowledge about introns and their regulatory elements, as well as understanding their functions. To achieve efficient expression, particularly in protein therapy, it is important to consider the use of either a truncated intron in a construct to recruit ISCSs (such as a promoter and enhancer), or a full-length intron in the appropriate position to facilitate processes involving SCEs, EJC, and IDREs. Alternatively, a strategy that combines both of these methods should be considered.

The main challenges in gene therapy are safety and efficiency ( 166 - 168 ). Therefore, both self-regulatory elements and potential regulatory elements derived from the introns are considered safe tools to address these issues ( 2 , 169 , 170 ). It has been revealed that the transposition of the L1 factor results in the activation of the p53-mediated apoptotic pathway in human cancer cells that have a wild-type p53 ( 171 ). So, IREs, especially L1 as a transgene, could act as a new approach for cancer therapy by stimulating apoptosis ( 171 ). Moreover, RNA reprogramming is an alternative method for gene therapy that involves correcting the sequence of specific transcripts through spliceosome-mediated RNA trans-splicing (SmaRT). This approach is applicable to genetic disorders caused by mutations in large genes or dominant-negative mutations ( 125 , 172 - 178 ).

4. Conclusions and Perspective

In summary, the intervening sequences of eukaryotic genes have critical functions in genomic regulation, and molecular evolution. In this regard, this paper briefly reviewed various known and potential functions of the intervening sequences. Based on various known and possible functions of the intervening sequences, as reviewed in this paper, introns not only carry information for splicing, but they also play many supportive roles in gene regulation at different levels. These roles are believed to function as useful tools in applications for various biological processes, particularly in the diagnosis and treatment of diseases. The effects of introns are related to splicing, SCEs (spliceosome catalytic elements), EJC (exon junction complex), IDREs (intron-derived regulatory elements), and ISCSs (intron splicing control sequences). Mutations in these elements that disrupt their functions play a significant role in disease development. Introns can contribute to numerous biological processes, including gene silencing, gene imprinting, transcription, mRNA metabolism, mRNA nuclear export, mRNA localization, mRNA surveillance, RNA editing, NMD, translation, protein stability, ribosome biogenesis, cell growth, embryonic development, apoptosis, molecular evolution, genome expansion, and proteome diversity through various mechanisms. Although the exact molecular mechanisms of some of the mentioned effects are determined, more basic research in this area is required. A major challenge is to unambiguously define the exact roles of each intron of a gene and their mechanisms on gene regulation. In general, having precise knowledge about these mechanisms may open new avenues for diagnosis and therapies against diseases, and it would also be advantageous in biotechnology applications.

Acknowledgments

We would like to gratefully thank the Ferdowsi University of Mashhad for its past and current support of research.

Ethics Approval and Consent to Participate

Not applicable.

Consent for publication

Not applicable.

Availability of data and materials

Data sharing is not applicable to this article as no datasets were generated during the current study.

Competing interest

The authors declare that they have no competing interests.

Funding

There is no funding for this study.

Authors’ contributions

A. H-M., M. M., and ME. TY. conceived of the existing idea, document extraction, and analysis. N.G-F. and A.B. draw the shapes and developed the contents. A. Z., M. Q., M. MM., and N. S. authors contributed to the final version of the manuscript. A. H-M. supervised the study. All authors discussed the consequences and contributed to the final manuscript.

References

  1. De Souza SJ, Long M, Gilbert W. Introns and gene evolution. Genes Cells. 1996; 1(6):493-505. DOI
  2. Kashima T, Rao N, Manley JL. An intronic element contributes to splicing repression in spinal muscular atrophy. Proc Natl Acad Sci U S A. 2007; 104(9):3426-3431. DOI
  3. Beard WA, Horton JK, Prasad R, Wilson SH. Eukaryotic base excision repair: new approaches shine light on mechanism. Annu Rev Biophys. 2019; 88:137-162. DOI
  4. Poverennaya I, Roytberg M. Spliceosomal introns: features, functions, and evolution. Biochemistry (Moscow).. 2020; 85 (7):725-734. DOI
  5. Herbert A, Rich A. RNA processing and the evolution of eukaryotes. Nat Genet. 1999; 21(3):265-269. DOI
  6. Bouaynaya N, Schonfeld D. The genomic structure: proof of the role of non-coding DNA. Conf Proc IEEE Eng Med Biol Soc. 2006; 1:4544-4547. DOI
  7. Rose AB. Introns as gene regulators: a brick on the accelerator. Frontiers in genetics. 2019; 9:672. DOI
  8. Palmiter RD, Sandgren EP, Avarbock MR, Allen DD, Brinster RL. Heterologous introns can enhance expression of transgenes in mice. Proc Natl Acad Sci U S A. 1991; 88(2):478-482. DOI
  9. Cottrell E, Maharaj A, Williams J, Chatterjee S, Cirillo G, Miraglia del Giudice E, et al. Growth Hormone Receptor (GHR) 6Ω Pseudoexon Activation: a Novel Cause of Severe Growth Hormone Insensitivity. J Clin Endocrinol Met. 2021. DOI
  10. Nesic D, Cheng J, Maquat LE. Sequences within the last intron function in RNA 3’-end formation in cultured cells. Mol Cell Biol. 1993; 13(6):3359-3369. DOI
  11. Antoniou M, Geraghty F, Hurst J, Grosveld F. Efficient 3’-end formation of human beta-globin mRNA in vivo requires sequences within the last intron but occurs independently of the splicing reaction. Nucleic Acids Res. 1998; 26(3):721-729. DOI
  12. Donath M, Mendel R, Cerff R, Martin W. Intron-dependent transient expression of the maize GapA1 gene. Plant Mol Biol. 1995; 28(4):667-676. DOI
  13. Jia J, Long Y, Zhang H, Li Z, Liu Z, Zhao Y, et al. Post- transcriptional splicing of nascent RNA contributes to widespread intron retention in plants. Nature Plants. 2020; 6(7):780-788. DOI
  14. Nott A, Meislin SH, Moore MJ. A quantitative analysis of intron effects on mammalian gene expression. RNA. 2003; 9(5):607-617. DOI
  15. Parenteau J, Maignon L, Berthoumieux M, Catala M, Gagnon V, Abou Elela S. Introns are mediators of cell response to starvation. Nature. 2019; 565(7741):612-617. DOI
  16. Liu H, Lyu HM, Zhu K, Van de Peer Y, Cheng ZM. The emergence and evolution of intron-poor and intronless genes in intron-rich plant gene families. The Plant Journal. 2021; 105(4):1072-1082. DOI
  17. Rohrer J, Conley ME. Transcriptional regulatory elements within the first intron of Bruton’s tyrosine kinase. Blood. 1998; 91(1):214-221. DOI
  18. Furger A, O’Sullivan JM, Binnie A, Lee BA, Proudfoot NJ. Promoter proximal splice sites enhance transcription. Genes Dev. 2002; 16(21):2792-2799. DOI
  19. Kim DS, Kim TH, Huh JW, Kim IC, Kim SW, Park HS, et al. Line Fusion Genes: a database of LINE expression in human genes. BMC Genomics. 2006; 7:139. DOI
  20. Eddy J, Maizels N. Conserved elements with potential to form polymorphic G-quadruplex structures in the first intron of human genes. Nucleic Acids Res. 2008; 36(4):1321-1333. DOI
  21. Tanaka Y, Asano T, Kanemitsu Y, Goto T, Yoshida Y, Yasuba K, et al. Positional differences of intronic transposons in pAMT affect the pungency level in chili pepper through altered splicing efficiency. The Plant Journal. 2019; 100(4):693-705. DOI
  22. Keilwagen J, Hartung F, Grau J. GeMoMa: homology-based gene prediction utilizing intron position conservation and RNA-seq data. Springer: Gene Prediction; 2019. DOI
  23. Le Hir H, Nott A, Moore MJ. How introns influence and enhance eukaryotic gene expression. Trends Biochem Sci. 2003; 28(4):215-220. DOI
  24. Desterro J, Bak-Gordon P, Carmo-Fonseca M. Targeting mRNA processing as an anticancer strategy. Nature Reviews Drug Discovery. 2020; 19(2):112-129. DOI
  25. Jiang W, Geng Y, Liu Y, Chen S, Cao S, Li W, et al. Genome-wide identification and characterization of SRO gene family in wheat: Molecular evolution and expression profiles during different stresses. Plant Physiology and Biochemistry. 2020; 154:590-611. DOI
  26. Valadkhan S. snRNAs as the catalysts of pre-mRNA splicing. Curr Opin Chem Biol. 2005; 9(6):603-608. DOI
  27. Joynt AT, Evans TA, Pellicore MJ, Davis-Marcisak EF, Aksit MA, Eastman AC, et al. Evaluation of both exonic and intronic variants for effects on RNA splicing allows for accurate assessment of the effectiveness of precision therapies. PLoS Gen. 2020; 16(10):e1009100. DOI
  28. Berget SM, Moore C, Sharp PA. Spliced segments at the 5’ terminus of adenovirus 2 late mRNA. Proc Natl Acad Sci U S A. 1977; 74(8):3171-3175. DOI
  29. Hua Y, Vickers TA, Okunola HL, Bennett CF, Krainer AR. Antisense masking of an hnRNP A1/A2 intronic splicing silencer corrects SMN2 splicing in transgenic mice. Am J Hum Genet. 2008; 82(4):834-848. DOI
  30. Borišek J, Casalino L, Saltalamacchia A, Mays SG, Malcovati L, Magistrato A. Atomic-Level Mechanism of Pre-mRNA Splicing in Health and Disease. Acc Chem Res. 2020; 54(1):144-154. DOI
  31. Majewski J, Ott J. Distribution and characterization of regulatory elements in the human genome. Genome Res. 2002; 12(12):1827-1836. DOI
  32. Faustino NA, Cooper TA. Pre-mRNA splicing and human disease. Genes Dev. 2003; 17(4):419-437. DOI
  33. Nilsen TW. The spliceosome: the most complex macromolecular machine in the cell?. Bioessays. 2003; 25(12):1147-1149. DOI
  34. Tang SJ, Shen H, An O, Hong H, Li J, Song Y, et al. Cis-and trans-regulations of pre-mRNA splicing by RNA editing enzymes influence cancer development. Nature Communicat. 2020; 11(1):1-17. DOI
  35. Tarn WY, Steitz JA. Pre-mRNA splicing: the discovery of a new spliceosome doubles the challenge. Trends Biochem Sci. 1997; 22(4):132-137. DOI
  36. Erkelenz S, Poschmann G, Ptok J, Müller L, Schaal H. Profiling of cis-and trans-acting factors supporting noncanonical splice site activation. RNA Biology. 2021; 18(1):118-130. DOI
  37. Tang SJ, Shen H, An O, Hong H, Li J, Song Y, et al. Cis- and trans-regulations of pre-mRNA splicing by RNA editing enzymes influence cancer development. Nature communicat. 2020; 11(1):799. DOI
  38. Moles-Fernández A, Domènech-Vivó J, Tenés A, Balmaña J, Diez O, Gutiérrez-Enríquez S. Role of splicing regulatory elements and in silico tools usage in the identification of deep intronic splicing variants in hereditary breast/ovarian cancer genes. Cancers. 2021; 13(13):3341. DOI
  39. Finke M, Brecht D, Stifel J, Gense K, Gamerdinger M, Hartig JS. Efficient splicing-based RNA regulators for tetracycline-inducible gene expression in human cell culture and C. elegans. Nucleic Acids Res. 2021. DOI
  40. Monteys AM, Hundley AA, Ranum PT, Tecedor L, Muehlmatt A, Lim E, et al. Regulated control of gene therapies by drug-induced splicing. Nature. 2021; 596(7871):291-295. DOI
  41. Haddad-Mashadrizeh A, Zomorodipour A, Izadpanah M, Sam MR, Ataei F, Sabouni F, et al. A systematic study of the function of the human beta-globin introns on the expression of the human coagulation factor IX in cultured Chinese hamster ovary cells. J Gene Med. 2009; 11(10):941-950. DOI
  42. Hahn S. Structure and mechanism of the RNA polymerase II transcription machinery. Nat Struct Mol Biol. 2004; 11(5):394-403. DOI
  43. Manley JL. Nuclear coupling: RNA processing reaches back to transcription. Nat Struct Biol. 2002; 9(11):790-791. DOI
  44. Kwek KY, Murphy S, Furger A, Thomas B, O’Gorman W, Kimura H, et al. U1 snRNA associates with TFIIH and regulates transcriptional initiation. Nat Struct Biol. 2002; 9(11):800-805. DOI
  45. O’Gorman W, Thomas B, Kwek KY, Furger A, Akoulitchev A. Analysis of U1 small nuclear RNA interaction with cyclin H. J Biol Chem. 2005; 280(44):36920-36925. DOI
  46. Tellier M, Maudlin I, Murphy S. Transcription and splicing: A two-way street. Wiley Interdisciplinary Reviews: RNA. 2020; 11(5):e1593. DOI
  47. Biswas J, Li W, Singer RH, Coleman RA. Imaging Organization of RNA Processing within the Nucleus. Cold Spring Harbor Perspectives in Biology. 2021;a039453. DOI
  48. Strasser K, Hurt E. Splicing factor Sub2p is required for nuclear mRNA export through its interaction with Yra1p. Nature. 2001; 413(6856):648-652. DOI
  49. Reed R, Hurt E. A conserved mRNA export machinery coupled to pre-mRNA splicing. Cell. 2002; 108(4):523-531. DOI
  50. Stewart M. Polyadenylation and nuclear export of mRNAs. J BiologChem. 2019; 294(9):2977-2987. DOI
  51. Schell T, Kulozik AE, Hentze MW. Integration of splicing, transport and translation to achieve mRNA quality control by the nonsense-mediated decay pathway. Genome Biol. 2002; 3(3):Reviews1006. DOI
  52. Tange TO, Nott A, Moore MJ. The ever-increasing comple-xities of the exon junction complex. Curr Opin Cell Biol. 2004; 16(3):279-284. DOI
  53. Le Hir H, Seraphin B. EJCs at the heart of translational control. Cell. 2008; 133(2):213-216. DOI
  54. Kwon OS, Mishra R, Safieddine A, Coleno E, Alasseur Q, Faucourt M, et al. Exon junction complex dependent mRNA localization is linked to centrosome organization during ciliogenesis. Nature communicat. 2021; 12(1):1-16. DOI
  55. Woodward LA, Mabin JW, Gangras P, Singh G. The exon junction complex: a lifelong guardian of mRNA fate. Wiley Interdisciplinary Reviews: RNA. 2017; 8(3):e1411. DOI
  56. Joseph B, Lai EC. The exon junction complex and intron removal prevent re-splicing of mRNA. PLoS Gen. 2021; 17 (5):e1009563. DOI
  57. Mabin JW, Woodward LA, Patton RD, Yi Z, Jia M, Wysocki VH, et al. The exon junction complex undergoes a compositional switch that alters mRNP structure and nonsense-mediated mRNA decay activity. Cell reports. 2018; 25(9):2431-2446. e7.
  58. Herold A, Suyama M, Rodrigues JP, Braun IC, Kutay U, Carmo-Fonseca M, et al. TAP (NXF1) belongs to a multigene family of putative RNA export factors with a conserved modular architecture. Mol Cell Biol. 2000; 20(23):8996-9008. DOI
  59. Clouse KN, Luo MJ, Zhou Z, Reed R. A Ran-independent pathway for export of spliced mRNA. Nat Cell Biol. 2001; 3(1):97-99. DOI
  60. Cheng C, Sharp PA. Regulation of CD44 alternative splicing by SRm160 and its potential role in tumor cell invasion. Mol Cell Biol. 2006; 26(1):362-370. DOI
  61. Hachet O, Ephrussi A. Drosophila Y14 shuttles to the posterior of the oocyte and is required for oskar mRNA transport. Curr Biol. 2001; 11(21):1666-1674. DOI
  62. Le Hir H, Gatfield D, Izaurralde E, Moore MJ. The exon-exon junction complex provides a binding platform for factors involved in mRNA export and nonsense-mediated mRNA decay. EMBO J. 2001; 20(17):4987-4997. DOI
  63. Shi H, Xu RM. Crystal structure of the Drosophila Mago nashi-Y14 complex. Genes Dev. 2003; 17(8):971-976. DOI
  64. Lykke-Andersen J, Shu MD, Steitz JA. Communication of the position of exon-exon junctions to the mRNA surveillance machinery by the protein RNPS1. Science. 2001; 293(5536):1836-1839. DOI
  65. Maquat LE. Nonsense-mediated mRNA decay: splicing, translation and mRNP dynamics. Nat Rev Mol Cell Biol. 2004; 5(2):89-99. DOI
  66. Makarova JA, Kramerov DA. Noncoding RNA of U87 host gene is associated with ribosomes and is relatively resistant to nonsense-mediated decay. Gene. 2005; 363:51-60. DOI
  67. Brogna S, Wen J. Nonsense-mediated mRNA decay (NMD) mechanisms. Nat Struct Mol Biol. 2009; 16(2):107-113. DOI
  68. Ying SY, Lin SL. Intron-derived microRNAs--fine tuning of gene functions. Gene. 2004; 342(1):25-28.
  69. Lin SL, Miller JD, Ying SY. Intronic microRNA (miRNA). J Biomed Biotechnol. 2006; 2006(4):26818. DOI
  70. Behl T, Kumar C, Makkar R, Gupta A, Sachdeva M. Intercalating the role of microRNAs in cancer: as enemy or protector. Asian Pacific journal of cancer prevention: APJCP. 2020; 21(3):593. DOI
  71. Esmailzadeh S, Mansoori B, Mohammadi A, Baradaran B. Regulatory roles of micro-RNAs in T cell autoimmunity. Immunological investigations. 2017; 46(8):864-879. DOI
  72. Hashemzadeh MR. Role of micro RNAs in stem cells, cardiac differentiation and cardiovascular diseases. Gene Reports. 2017; 8:11-6. DOI
  73. Stark A, Brennecke J, Russell RB, Cohen SM. Identification of Drosophila MicroRNA targets. PLoS Biol. 2003; 1(3):E60. DOI
  74. Pederson T. RNA interference and mRNA silencing, 2004: how far will they reach?. Mol Biol Cell. 2004; 15(2):407-410. DOI
  75. Tomasello L, Distefano R, Nigita G, Croce CM. The microRNA family gets wider: the isomiRs classification and role. Frontiers in Cell and Developmental Biology. 2021; 9DOI
  76. Ambros V. MicroRNA pathways in flies and worms: growth, death, fat, stress, and timing. Cell. 2003; 113(6):673-676. DOI
  77. Islam ABMM, Mohammad E, Khan M. Aberration of the modulatory functions of intronic microRNA hsa-miR-933 on its host gene ATF2 results in type II diabetes mellitus and neurodegenerative disease development. Human Genomics. 2020; 14(1):1-11. DOI
  78. Palatnik JF, Allen E, Wu X, Schommer C, Schwab R, Carrin-gton JC, et al. Control of leaf morphogenesis by microRNAs. Nature. 2003; 425(6955):257-263. DOI
  79. Cao X, Fan Q-L. LncRNA MIR503HG promotes high-glucose-induced proximal tubular cell apoptosis by targeting miR-503-5p/bcl-2 pathway. Diabetes, Metabolic Syndrome and Obesity: Targets and Therapy. 2020; 13:4507. DOI
  80. Qu LH, Henras A, Lu YJ, Zhou H, Zhou WX, Zhu YQ, et al. Seven novel methylation guide small nucleolar RNAs are processed from a common polycistronic transcript by Rat1p and RNase III in yeast. Mol Cell Biol. 1999; 19(2):1144-1158. DOI
  81. Bachellerie JP, Cavaille J, Huttenhofer A. The expanding sno- RNA world. Biochimie. 2002; 84(8):775-790. DOI
  82. Frazier MN, Pillon MC, Kocaman S, Gordon J, Stanley RE. Structural overview of macromolecular machines involved in ribosome biogenesis. Current Opinion in Structural Biology. 2021; 67:51-60. DOI
  83. Kumar V. Ribosomal biogenesis in eukaryotes. Elsevier: Emerging Con-cepts in Ribosome Structure, Biogenesis, and Function; 2021. DOI
  84. Bratkovič T, Božič J, Rogelj B. Functional diversity of small nucleolar RNAs. Nucleic acids research. 2020; 48(4):1627-1651. DOI
  85. Huttenhofer A, Kiefmann M, Meier-Ewert S, O’Brien J, Lehrach H, Bachellerie JP, et al. RNomics: an experimental approach that identifies 201 candidates for novel, small, non-messenger RNAs in mouse. EMBO J. 2001; 20(11):2943-2953. DOI
  86. Reinhart BJ, Bartel DP. Small RNAs correspond to centromere heterochromatic repeats. Science. 2002; 297(5588):1831. DOI
  87. Tritto P, Specchia V, Fanti L, Berloco M, D’Alessandro R, Pimpinelli S, et al. Structure, regulation and evolution of the crystal-Stellate system of Drosophila. Genetica. 2003; 117(2-3):247-257. DOI
  88. Matzke M, Aufsatz W, Kanno T, Daxinger L, Papp I, Mette MF, et al. Genetic analysis of RNA-mediated transcriptional gene silencing. Biochim Biophys Acta. 2004; 1677(1-3):129-141. DOI
  89. Rinn JL, Kertesz M, Wang JK, Squazzo SL, Xu X, Brugmann SA, et al. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell. 2007; 129(7):1311-1323. DOI
  90. Polavarapu N, Marino-Ramirez L, Landsman D, McDonald JF, Jordan IK. Evolutionary rates and patterns for human transcription factor binding sites derived from repetitive DNA. BMC Genomics. 2008; 9:226. DOI
  91. Parris GE. Developmental diseases and the hypothetical Master Development Program. Med Hypotheses. 2010; 74(3):564-573. DOI
  92. Qian W, Zhang J. Codon usage bias and nuclear mRNA concentration: Correlation vs. causation. Proceedings of the National Academy of Sciences. 2021; 118(20)DOI
  93. Darzacq X, Jady BE, Verheggen C, Kiss AM, Bertrand E, Kiss T. Cajal body-specific small nuclear RNAs: a novel class of 2’-O-methylation and pseudouridylation guide RNAs. EMBO J. 2002; 21(11):2746-2756. DOI
  94. Nishihara H, Smit AF, Okada N. Functional noncoding sequences derived from SINEs in the mammalian genome. Genome Res. 2006; 16(7):864-874. DOI
  95. Huppert JL. Hunting G-quadruplexes. Biochimie. 2008; 90(8): 1140-1148. DOI
  96. Vagner S, Vagner C, Mattaj IW. The carboxyl terminus of vertebrate poly(A) polymerase interacts with U2AF 65 to couple 3’-end processing and splicing. Genes Dev. 2000; 14 (4):403-413. DOI
  97. Fong YW, Zhou Q. Stimulatory effect of splicing factors on transcriptional elongation. Nature. 2001; 414(6866):929-933. DOI
  98. Polak P, Domany E. Alu elements contain many binding sites for transcription factors and may play a role in regulation of developmental processes. BMC Genomics. 2006; 7:133. DOI
  99. Tahiliani J, Leisk J, Aradhya K, Ouyang K, Aradhya S, Nykamp K. Utility of RNA Sequencing Analysis in the Context of Genetic Testing. Current Genetic Medicine Reports. 2020;1-7. DOI
  100. Duquette ML, Handa P, Vincent JA, Taylor AF, Maizels N. Intracellular transcription of G-rich DNAs induces formation of G-loops, novel structures containing G4 DNA. Genes Dev. 2004; 18(13):1618-1629. DOI
  101. Duquette ML, Pham P, Goodman MF, Maizels N. AID binds to transcription-induced structures in c-MYC that map to regions associated with translocation and hypermutation. Oncogene. 2005; 24(38):5791-5798. DOI
  102. Duquette ML, Huber MD, Maizels N. G-rich proto-oncogenes are targeted for genomic instability in B-cell lymphomas. Cancer Res. 2007; 67(6):2586-2594. DOI
  103. Larson ED, Duquette ML, Cummings WJ, Streiff RJ, Maizels N. MutSalpha binds to and promotes synapsis of transcriptionally activated immunoglobulin switch regions. Curr Biol. 2005; 15(5):470-474. DOI
  104. Burge S, Parkinson GN, Hazel P, Todd AK, Neidle S. Quadruplex DNA: sequence, topology and structure. Nucleic Acids Res. 2006; 34(19):5402-5415. DOI
  105. Maizels N. Dynamic roles for G4 DNA in the biology of eukaryotic cells. Nat Struct Mol Biol. 2006; 13(12):1055-1059. DOI
  106. Phan AT, Kuryavyi V, Patel DJ. DNA architecture: from G to Z. Curr Opin Struct Biol. 2006; 16(3):288-298. DOI
  107. Yang H, Zhou Y, Liu J. G-quadruplex DNA for construction of biosensors. TrAC Trends in Analytical Chemistry. 2020; 116060. DOI
  108. Huppert JL, Balasubramanian S. Prevalence of quadruplexes in the human genome. Nucleic Acids Res. 2005; 33(9):2908-2916. DOI
  109. Todd AK, Johnston M, Neidle S. Highly prevalent putative quadruplex sequence motifs in human DNA. Nucleic Acids Res. 2005; 33(9):2901-7. DOI
  110. Du Z, Kong P, Gao Y, Li N. Enrichment of G4 DNA motif in transcriptional regulatory region of chicken genome. Biochem Biophys Res Commun. 2007; 354(4):1067-1070. DOI
  111. Huppert JL, Balasubramanian S. G-quadruplexes in promoters throughout the human genome. Nucleic Acids Res. 2007; 35(2): 406-413. DOI
  112. Zhao Y, Du Z, Li N. Extensive selection for the enrichment of G4 DNA motifs in transcriptional regulatory regions of warm blooded animals. FEBS Lett. 2007; 581(10):1951-1956. DOI
  113. Menon S, Piramanayakam S, Agarwal G. Computational iden-tification of promoter regions in prokaryotes and Eukaryotes. EPRA International Journal of Agriculture and Rural Economic Research (ARER).. 2021; 9(7):21-28. DOI
  114. Haddad-Mashadrizeh A, Hemmat J, Aslamkhan M. Intronic regions of the human coagulation factor VIII gene harboring transcription factor binding sites with a strong bias towards the short-interspersed elements. Heliyon. 2020; 6(9):e04727. DOI
  115. Gehring NH, Roignant J-Y. Anything but ordinary–emerging splicing mechanisms in eukaryotic gene regulation. Trends in Genetics. 2020. DOI
  116. Petibon C, Malik Ghulam M, Catala M, Abou Elela S. Regulation of ribosomal protein genes: An ordered anarchy. Wiley Interdisciplinary Reviews RNA. 2021; 12(3):e1632. DOI
  117. Lopez AJ. Alternative splicing of pre-mRNA: developmental consequences and mechanisms of regulation. Annu Rev Genet. 1998; 32:279-305. DOI
  118. Modrek B, Lee C. A genomic view of alternative splicing. Nat Genet. 2002; 30(1):13-19. DOI
  119. Bruno IG, Jin W, Cote GJ. Correction of aberrant FGFR1 alternative RNA splicing through targeting of intronic regulatory elements. Hum Mol Genet. 2004; 13(20):2409-2420. DOI
  120. Ladomery MR, Harper SJ, Bates DO. Alternative splicing in angiogenesis: the vascular endothelial growth factor paradigm. Cancer Lett. 2007; 249(2):133-142. DOI
  121. Lamaa A, Humbert J, Aguirrebengoa M, Cheng X, Nicolas E, Côté J, et al. Integrated analysis of H2A. Z isoforms function reveals a complex interplay in gene regulation. Elife. 2020; 9:e5337-5. DOI
  122. Sun H, Chasin LA. Multiple splicing defects in an intronic false exon. Mol Cell Biol. 2000; 20(17):6414-6425. DOI
  123. Vela E, Roca X, Isamat M. Identification of novel splice variants of the human CD44 gene. Biochem Biophys Res Commun. 2006; 343(1):167-170. DOI
  124. Di Segni G, Gastaldi S, Tocchini-Valentini GP. Cis- and trans-splicing of mRNAs mediated by tRNA sequences in eukaryotic cells. Proc Natl Acad Sci U S A. 2008; 105(19):6864-6869. DOI
  125. Viles KD, Sullenger BA. Proximity-dependent and proximity-independent trans-splicing in mammalian cells. RNA. 2008; 14(6):1081-1094. DOI
  126. Hasler J, Strub K. Alu elements as regulators of gene expression. Nucleic Acids Res. 2006; 34(19):5491-5497. DOI
  127. Bhadra M, Howell P, Dutta S, Heintz C, Mair WB. Alterna-tive splicing in aging and longevity. Human genetics. 2020; 139 (3):357-369. DOI
  128. Sorek R, Ast G, Graur D. Alu-containing exons are alternatively spliced. Genome Res. 2002; 12(7):1060-1067. DOI
  129. Pérez-Molina R, Arzate-Mejía RG, Ayala-Ortega E, Guerrero G, Meier K, Suaste-Olmos F, et al. An intronic Alu element attenuates the transcription of a long non-coding RNA in human cell lines. Frontiers In Genetics. 2020; 11:928. DOI
  130. Lozano G, Francisco-Velilla R, Martinez-Salas E. Deconstruc-ting internal ribosome entry site elements: an update of structural motifs and functional divergences. Royal Society Open Biology. 2018; 8(11):180155. DOI
  131. Babich V, Aksenov N, Alexeenko V, Oei SL, Buchlow G, Tomilin N. Association of some potential hormone response elements in human genes with the Alu family repeats. Gene. 1999; 239(2):341-349. DOI
  132. Li W, Kuzoff R, Wong CK, Tucker A, Lynch M. Characterization of newly gained introns in Daphnia populations. Genome biology and evolution. 2014; 6(9):2218-2234. DOI
  133. Makalowski W. Genomic scrap yard: how genomes utilize all that junk. Gene. 2000; 259(1-2):61-67. DOI
  134. Nekrutenko A, Li WH. Transposable elements are found in a large number of human protein-coding genes. Trends Genet. 2001; 17(11):619-621. DOI
  135. Corley M, Flynn RA, Lee B, Blue SM, Chang HY, Yeo GW. Footprinting SHAPE-eCLIP Reveals Transcriptome-wide Hydrogen Bonds at RNA-Protein Interfaces. Molecular Cell. 2020; 80(5):903-914. e8.
  136. Jo B-S, Choi SS. Introns: the functional benefits of introns in genomes. Genomics & informatics. 2015; 13(4):112. DOI
  137. Baralle FE, Giudice J. Alternative splicing as a regulator of development and tissue identity. Nature Rev Molr cell biolog. 2017; 18(7):437-451. DOI
  138. Chang YF, Imam JS, Wilkinson MF. The nonsense-mediated decay RNA surveillance pathway. Annu Rev Biochem. 2007; 76:51-74. DOI
  139. Hagiwara M. Alternative splicing: a new drug target of the post-genome era. Biochim Biophys Acta. 2005; 1754(1-2):324-331. DOI
  140. Wei C, Xie W, Huang X, Mo X, Liu Z, Wu G, et al. Profiles of alternative splicing events in the diagnosis and prognosis of Gastric Cancer. J Cancer. 2021; 12(10):2982. DOI
  141. Eblen ST. Extracellular-regulated kinases: signaling from Ras to ERK substrates to control biological outcomes. Adv Cancer Res. 2018; 138:99-142. DOI
  142. Hujová P, Souček P, Grodecká L, Grombiříková H, Ravčuková B, Kuklínek P, et al. Deep intronic mutation in SERPING1 caused hereditary angioedema through pseudoexon activation. Journal of clinical immunology. 2020; 40(3):435-46. DOI
  143. Venables JP. Aberrant and alternative splicing in cancer. Cancer Res. 2004; 64(21):7647-5764. DOI
  144. Venables JP. Unbalanced alternative splicing and its significance in cancer. Bioessays. 2006; 28(4):378-386. DOI
  145. Rhine CL, Cygan KJ, Soemedi R, Maguire S, Murray MF, Monaghan SF, et al. Hereditary cancer genes are highly susceptible to splicing mutations. PLoS Gen. 2018; 14(3):e1007231. DOI
  146. Kashkan I, Timofeyenko K, Kollárová E, Růžička K. In Vivo Reporters for Visualizing Alternative Splicing of Hormonal Genes. Plants. 2020; 9(7):868. DOI
  147. Biamonti G, Infantino L, Gaglio D, Amato A. An intricate connection between alternative splicing and phenotypic plasticity in development and cancer. Cells. 2020; 9(1):34. DOI
  148. Sneath RJ, Mangham DC. The normal structure and function of CD44 and its role in neoplasia. Mol Pathol. 1998; 51(4):191-200. DOI
  149. Chalfant CE, Rathman K, Pinkerman RL, Wood RE, Obeid LM, Ogretmen B, et al. De novo ceramide regulates the alternative splicing of caspase 9 and Bcl-x in A549 lung adenocarcinoma cells. Dependence on protein phosphatase-1. J Biol Chem. 2002; 277(15):12587-12595. DOI
  150. Makhafola TJ, Mbele M, Yacqub-Usman K, Hendren A, Haigh DB, Blackley Z, et al. Apoptosis in cancer cells is induced by alternative splicing of hnRNPA2/B1 through splicing of Bcl-x, a mechanism that can be stimulated by an extract of the South African Medicinal Plant, Cotyledon orbiculata. Frontiers in Oncology. 2020; 10DOI
  151. Blake D, Lynch KW. The three as: Alternative splicing, alternative polyadenylation and their impact on apoptosis in immune function. Immunol Rev. 2021. DOI
  152. López-Martínez A, Soblechero-Martín P, de-la-Puente-Ovejero L, Nogales-Gadea G, Arechavala-Gomeza V. An overview of alternative splicing defects implicated in myotonic dystrophy type i. Genes. 2020; 11(9):1109. DOI
  153. Hofmann Y, Lorson CL, Stamm S, Androphy EJ, Wirth B. Htra2-beta 1 stimulates an exonic splicing enhancer and can restore full-length SMN expression to survival motor neuron 2 (SMN2). Proc Natl Acad Sci U S A. 2000; 97(17):9618-9623. DOI
  154. Nissim-Rafinia M, Chiba-Falek O, Sharon G, Boss A, Kerem B. Cellular and viral splicing factors can modify the splicing pattern of CFTR transcripts carrying splicing mutations. Hum Mol Genet. 2000; 9(12):1771-1778. DOI
  155. Helman G, Takanohashi A, Hagemann TL, Perng MD, Walkiewicz M, Woidill S, et al. Type II Alexander disease caused by splicing errors and aberrant overexpression of an uncharacterized GFAP isoform. Human mutation. 2020; 41(6):1131-1137. DOI
  156. Sazani P, Kole R. Modulation of alternative splicing by antisense oligonucleotides. Prog Mol Subcell Biol. 2003; 31:217-239. DOI
  157. Celotto AM, Lee JW, Graveley BR. Exon-specific RNA interference: a tool to determine the functional relevance of proteins encoded by alternatively spliced mRNAs. Methods Mol Biol. 2005; 309:273-282. DOI
  158. Scharner J, Ma WK, Zhang Q, Lin K-T, Rigo F, Bennett CF, et al. Hybridization-mediated off-target effects of splice-switching antisense oligonucleotides. Nucleic Acids Res. 2020; 48(2):802-816. DOI
  159. Halloy F, Iyer PS, Ćwiek P, Ghidini A, Barman-Aksözen J, Wildner-Verhey van Wijk N, et al. Delivery of oligonucleotides to bone marrow to modulate ferrochelatase splicing in a mouse model of erythropoietic protoporphyria. Nucleic Acids Res. 2020; 48(9):4658-71. DOI
  160. Pilch B, Allemand E, Facompre M, Bailly C, Riou JF, Soret J, et al. Specific inhibition of serine- and arginine-rich splicing factors phosphorylation, spliceosome assembly, and splicing by the antitumor drug NB-506. Cancer Res. 2001; 61(18):6876-6884.
  161. Chen Y, Huang M, Liu X, Huang Y, Liu C, Zhu J, et al. Alternative splicing of mRNA in colorectal cancer: new strategies for tumor diagnosis and treatment. Cell Death & Disease. 2021; 12(8):1-16. DOI
  162. Varani L, Spillantini MG, Goedert M, Varani G. Structural basis for recognition of the RNA major groove in the tau exon 10 splicing regulatory element by aminoglycoside antibiotics. Nucleic Acids Res. 2000; 28(3):710-719. DOI
  163. Liu X, Jiang Q, Mansfield SG, Puttaraju M, Zhang Y, Zhou W, et al. Partial correction of endogenous DeltaF508 CFTR in human cystic fibrosis airway epithelia by spliceosome-mediated RNA trans-splicing. Nat Biotechnol. 2002; 20(1):47-52. DOI
  164. Lu S, Cullen BR. Analysis of the stimulatory effect of splicing on mRNA production and utilization in mammalian cells. RNA. 2003; 9(5):618-630. DOI
  165. Sam MR, Zomorodipour A, Shokrgozar MA, Ataei F, Haddad-Mashadrizeh A, Amanzadeh A. Enhancement of the human factor IX expression, mediated by an intron derived fragment from the rat aldolase B gene in cultured hepatoma cells. Biotechnol Lett. 2010; 32(10):1385-1392. DOI
  166. Appledorn DM, Patial S, McBride A, Godbehere S, Van Rooijen N, Parameswaran N, et al. Adenovirus vector-induced innate inflammatory mediators, MAPK signaling, as well as adaptive immune responses are dependent upon both TLR2 and TLR9 in vivo. J Immunol. 2008; 181(3):2134-2144. DOI
  167. Tang R, Xu Z. Gene therapy: A double-edged sword with great powers. Mol Cell Biochem. 2020; 474(1):73-81. DOI
  168. Jiao Y, Xia ZL, Ze LJ, Jing H, Xin B, Fu S. Research Progress of nucleic acid delivery vectors for gene therapy. Biomedical microdevices. 2020; 22(1):1-10. DOI
  169. Schuppe HC, Meinhardt A. Immune privilege and inflammation of the testis. Chem Immunol Allergy. 2005; 88:1-14. DOI
  170. Willerth SM, Sakiyama-Elbert SE. Combining stem cells and biomaterial scaffolds for constructing tissues and cell delivery. 2008.
  171. Haoudi A, Semmes OJ, Mason JM, Cannon RE. Retrotrans-position-Competent Human LINE-1 Induces Apoptosis in Cancer Cells With Intact p53. J Biomed Biotechnol. 2004; 2004(4):185-194.
  172. Yang Y, Walsh CE. Spliceosome-mediated RNA trans-splicing. Mol Ther. 2005; 12(6):1006-1012. DOI
  173. Chao H, Walsh CE. RNA repair for haemophilia A. Expert Rev Mol Med. 2006; 8(1):1-8. DOI
  174. Wood M, Yin H, McClorey G. Modulating the expression of disease genes with RNA-based therapy. PLoS Genet. 2007; 3(6):e109. DOI
  175. Wang J, Mansfield SG, Cote CA, Jiang PD, Weng K, Amar MJ, et al. Trans-splicing into highly abundant albumin transcripts for production of therapeutic proteins in vivo. Mol Ther. 2009; 17(2):343-351. DOI
  176. To TK, Nishizawa Y, Inagaki S, Tarutani Y, Tominaga S, Toyoda A, et al. RNA interference-independent reprogramming of DNA methylation in Arabidopsis. Nature Plants. 2020; 6(12):1455-1467. DOI
  177. Hong EM, Ingemarsdotter CK, Lever AM. Therapeutic applications of trans-splicing. British Medical Bulletin. 2020; 136(1):4-20. DOI
  178. Riedmayr LM. SMaRT for therapeutic purposes. Springer: Chimeric RNA; 2020. DOI
  179. Luo M-J, Zhou Z, Magni K, Christoforides C, Rappsilber J, Mann M, et al. Pre-mRNA splicing and mRNA export linked by direct interactions between UAP56 and Aly. Nature. 2001; 413(6856):644-647. DOI
  180. Anderson CM, Kohorn BD. Inactivation of Arabidopsis SIP1 leads to reduced levels of sugars and drought tolerance. J Plant Physiolog. 2001; 158(9):1215-1219. DOI
  181. Besse F, Ephrussi A. Translational control of localized mRNAs: restricting protein synthesis in space and time. Nat Rev Mol Cell Biol. 2008; 9(12):971-980. DOI
  182. Gustafsson C, Reid R, Greene PJ, Santi DV. Identification of new RNA modifying enzymes by iterative genome search using known modifying enzymes as probes. Nucleic Acids Res. 1996; 24(19):3756-3762. DOI
  183. Liu J, Perumal NB, Oldfield CJ, Su EW, Uversky VN, Dunker AK. Intrinsic disorder in transcription factors. Biochemistry. 2006; 45(22):6873-6888. DOI
  184. Cooke C, Hans H, Alwine JC. Utilization of splicing elements and polyadenylation signal elements in the coupling of polyadenylation and last-intron removal. Mol Cell Biol. 1999; 19(7):4971-4979. DOI